Photocatalytic N-Doped Tio For Self-Cleaning of Limestones: T E P J P

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Eur. Phys. J.

Plus (2019) 134: 539


DOI 10.1140/epjp/i2019-12981-6
THE EUROPEAN
PHYSICAL JOURNAL PLUS
Regular Article

Photocatalytic N-doped TiO2 for self-cleaning of limestones 

Laura Fornasini1,a , Laura Bergamonti2 , Federica Bondioli3 , Danilo Bersani1 , Laura Lazzarini4 , Yaron Paz5 , and
Pier Paolo Lottici1
1
Department of Mathematical, Physical and Computer Sciences, University of Parma, Parco Area delle Scienze 7/a, 43124
Parma, Italy
2
Department of Chemistry, Life Sciences and Environmental Sustainability, University of Parma, Parco Area delle Scienze
17/a, 43124 Parma, Italy
3
Department of Applied Science and Technology, Politecnico di Torino, Corso Duca degli Abruzzi, 24, 10129 Torino, Italia
4
IMEM-CNR Parco Area delle Scienze 37/A - 43124 Parma, Italy
5
Department of Chemical Engineering, Technion-Israel Institute of Technology, Haifa, 32000, Israel

Received: 27 September 2018 / Revised: 22 July 2019


Published online: 31 October 2019

c Società Italiana di Fisica / Springer-Verlag GmbH Germany, part of Springer Nature, 2019

Abstract. Nanocrystalline nitrogen-doped TiO2 photocatalysts were applied as surface coatings to three
typical Israeli limestones, Maccabim, Halila and Hebron, widely used for historic buildings. Two different
N-doped TiO2 sols were synthesised by sol-gel method starting from titanium oxysulfate (TiOSO4 ) as
Ti precursor: a new N-doped TiO2 coating (N-TiF), obtained in neutral conditions, was developed and
compared to a N-doped TiO2 (N-TiA), obtained in acidic medium. XRD and Raman spectroscopy confirm
the nanocrystalline nature of N-doped TiO2 , mainly in anatase phase, with size ∼ 4 nm as revealed by
TEM analysis. Diffuse Reflectance Spectroscopy indicates a red-shift in the absorption spectra for both
nanopowders, even if the X-ray Photoelectron spectroscopy results are not conclusive on the state and
location of the N doping ions. The self-cleaning efficiency of the two TiO2 coatings was assessed on the three
Israeli carbonate stones and compared to commercial TiO2 (Evonik P25) by the photoinduced degradation
of two organic dyes (Methyl orange and Rhodamine B) and by following hydrophilicity changes, through
static contact angle measurements. The harmlessness of these coatings was confirmed by colorimetric and
water capillary absorption measurements.

1 Introduction
Nanomaterials are currently investigated as coating materials for consolidating and protecting stones of monumental
and archaeological interest. In particular, the photocatalytic self-cleaning properties of innovative materials represent a
great promise in the decomposition and removal of soiling compounds, strongly required to slow down the degradation
processes and to preserve the original appearance of artefacts [1–3]. Photocatalysis is referred to the acceleration of a
reaction occurring when a material, typically a semiconductor, absorbs light of appropriate energy, inducing the charge
separation of electrons and holes, producing reactive oxidizing species and leading to the transformation of pollutants
into harmless species [4]. Nanomaterials are investigated for their high surface-to-volume ratio with large density of
active sites available for adsorption and catalysis, enhancing the chemical reactivity and photocatalytic efficiency. The
experimental synthesis parameters are important for the suitability and the effectiveness of the treatments: they can
influence and modify the morphology, the particle sizes and the crystalline structures of the nanomaterials [5].
Titanium dioxide in nanocrystalline form, having high reactivity and chemical stability, is considered the most
promising photocatalyst, to be used for decomposition of environmental pollutants [6,7] and removal of stains deposited
on the outdoor surface of historical buildings and Cultural Heritage artefacts [8–13]. To obtain nanocrystalline TiO2
with high purity in absence of unwanted by-products, the sol-gel method is widely used, working at relatively low
temperatures and enabling the control of the crystallite sizes: nanocrystals in the range 5–10 nm are easily obtained,

Focus Point on “Scientific Research in Conservation Science” edited by L. Bellot-Gurlet, D. Bersani, P. Vandenabeele.

Supplementary material in the form of a .pdf file available from the Journal web page at
https://doi.org/10.1140/epjp/i2019-12981-6
a
e-mail: laura.fornasini@studenti.unipr.it
Page 2 of 13 Eur. Phys. J. Plus (2019) 134: 539

these sizes being generally suggested to improve the photocatalytic activity [14]. On the other hand, the band gap
of TiO2 requires UV light to promote photocatalytic processes and this limits to about 5% of the available solar
irradiation. The goal to produce a visible light active photocatalyst can be achieved through doping with impurities
(cations or anions) at the lattice/interstitial sites which leads to the formation of localised energy levels (donor
or acceptor levels) [15]. The dopant choice is based on ionic radius, oxidation state and electronic configuration.
Nitrogen atoms can be easily hosted in the TiO2 structure, having atomic size comparable with oxygen. N-doped
TiO2 (N-TiO2 ) can show a shift of the absorption edge into the visible (from 3.2 eV up to about 2.3 eV), enhancing
the photocatalytic activity upon solar irradiation. For the synthesis of N-TiO2 , the sol-gel route, a very versatile
technique, allows the growth of TiO2 nanoparticles and their N doping through the hydrolysis of titanium precursors
in presence of nitrogen sources. Typically, titanium alkoxides (titanium tetraisopropoxide, tetrabutyl orthotitanate) or
titanium salts (titanium oxysulfate, titanium tetrachloride) have been used. Nitrogen sources include amines, nitrates,
urea and ammonium salts [4].
In most cases, titania nanosols are synthesised in acidic conditions and their efficiency for photocatalytic degra-
dation of pollutants is well established [10,16]. However, their use on carbonate stones may be harmful. Obtaining
coatings in acidic medium may require additional procedures before application of the coatings on soft stones, like
biocalcarenites and carbonate stones. For this reason, in this study we focused on the development of a new N-doped
TiO2 coating, obtained in neutral conditions, and on the evaluation of its photocatalytic efficiency and harmlessness
when applied on carbonate stone samples. The new neutral coating is compared to a N-doped TiO2 (a product hence-
forth called N-TiA), obtained in acidic medium, whose performance has been reported in a previous work [17]. The
efficient improvement of the N-doped TiO2 coating under visible irradiation and the applicability for self-cleaning
on travertine stone was demonstrated, still requiring additional washing and aqueous dispersion to be suitable for
application on carbonate stones.
Here we focus on the synthesis and characterisation of the neutral nitrogen doped TiO2 photocatalyst, called N-
TiF, obtained using NH4 OH and NH4 F as N sources instead of NH4 OH and HNO3 , which were used in the synthesis
of N-TiA. The self-cleaning properties of the N-TiF coating are here compared to those of the N-TiA product and of
the commercial Aeroxide TiO2 P25 by Evonik, applied on three typical Israeli limestones. The photoinduced changes
in stone wettability and the photodegradation of two synthetic organic dyes (Methyl orange and Rhodamine B) under
UV-Vis irradiation have been measured.
To ensure that the treatments do not alter the original characteristics of the stones, measurements of water
absorption by capillarity were performed and colour changes produced by the coating deposition were evaluated.

2 Material and methods


2.1 Synthesis of N-doped TiO2 nanoparticles

Chemicals and materials used for the synthesis were titanium(IV) oxysulfate (TiOSO4 , Aldrich), nitric acid (HNO3 ,
65%, Carlo Erba), ammonium hydroxide (NH4 OH, 26%, Carlo Erba) and ammonium fluoride (NH4 F): all reagents
and Aeroxide P25 (Evonik) were used as received.
Two different N-doped TiO2 sols (0.1M), N-TiA and N-TiF, were obtained. Both TiO2 nano-sols were synthesised
starting from 0.01 mole of titanium(IV) oxysulfate, TiOSO4 , hydrolysed in 100 ml distilled water at 40 ◦ C. After
cooling, NH4 OH was slowly added dropwise to the solution, up to pH ≈ 7, forming a white precipitate that was
separated by centrifugation and repeatedly washed with distilled water, until residual SO4 2− ions (as confirmed by
BaCl2 test) could no longer be observed. The resulting product was suspended in 100 ml distilled water, followed
by peptization with HNO3 until pH ≈ 1.5 for N-TiA sample [17], or addition of NH4 F in the molar proportion 1:3
(TiOSO4 :NH4 F), to obtain N-TiF sample. After vigorous stirring at room temperature for 30 minutes, both sols were
refluxed at 100 ◦ C for 2 hours to induce crystallization, thus obtaining the final nano-sol. The final pH was 1.6 and
7.3 for N-TiA and N-TiF, respectively.

2.2 Characterisation of N-doped TiO2 samples

Small amounts of the sols were centrifuged at 1000 rpm, repeatedly washed and finally dried at 30 ◦ C for 24 h. Room-
temperature X-ray diffraction (XRD) patterns were collected in 2θ range of 10◦ –60◦ by a Thermo ARL X’TRA X-ray
diffractometer with Si-Li detector, using Cu Kα radiation at 40 kV and 40 mA, at 0.2◦ step size, count time 10 s, on
the powders spread on a conventional glass sample holder. Powdered crystalline silicon was used for 2θ calibration.
XRD patterns were compared with standards of anatase, brookite and rutile diffractograms (JCPDS cards #21–
1272, #29–1360 and #76–1940, respectively). The average diameter of the crystallites was estimated by the Scherrer
equation [16]. Transmission electron microscopy (TEM) observations, were carried out by a high-resolution (0.18 nm)
field-emission JEOL 2200FS microscope, equipped with in-column Ω energy filter, 2 high-angle annular dark field
Eur. Phys. J. Plus (2019) 134: 539 Page 3 of 13

(HAADF) detectors, and X-ray microanalysis (EDS). The samples for the observation were prepared by dispersing a few
drops of the nanoparticles TiO2 solution, appropriately diluted, onto 300-mesh carbon coated copper grids and drying
them in air. Raman scattering measurements were performed using a Horiba-Jobin Yvon LabRAM micro-spectrometer.
Spectra were recorded at room temperature with the 632.8 nm line of a He–Ne laser at a spectral resolution of about 1.5–
2 cm−1 . The Rayleigh radiation was blocked by an edge filter and the Raman light was dispersed by an 1800 grooves/mm
holographic grating on a 1024/256 pixels CCD, Peltier cooled. The entrance-slit width was fixed at 100 μm. The laser
power on the sample was adjusted at < 2 mW by density filters to avoid unacceptable thermal effects. Spectra were
collected using a ×50 microscope objective. The wavenumber calibration was obtained by the 520.6 cm−1 Raman band
of silicon. Typical exposures were 10–60 s, repeated three to five times and the data analysis was performed by LabSpec
built-in software. UV-Vis diffuse reflectance spectroscopy (DRS) measurements on TiO2 powdered samples were carried
out by a Varian 2390 spectrophotometer equipped with an integrating sphere. The spectra were recorded in reflectance
mode at room temperature, with BaSO4 as reference, scanning in the 290–800 nm wavelength range. Absorption
coefficients and energy gaps were calculated by means of a Kubelka-Munk treatment [18] and Tauc plots [19]. TiO2
samples were investigated by X-ray photoelectron spectroscopy (XPS) with Al Kα source (hν = 1486.6 eV) using
a Thermo Scientific Sigma Probe equipped with hemispherical energy analyser. The powders were attached onto a
carbon tape. The signal from the C1s peak at 285 eV was taken as internal control, to account for charging effects.
The samples were first scanned in fast mode, where the signal is collected in a wide angle (from 24◦ to 84◦ relative to
the sample normal), then specific features were finely scanned in a slower mode collecting the signal from a narrower
window (from 24◦ to 38◦ degrees relative to the normal). This procedure gives a bulk-emphasizing measurement. The
survey scanning was obtained with a step size of 1 eV, whereas specific peaks were analysed at a step size of 0.05 eV. The
lateral size of the beam was 200 μm by 400 μm, which is large enough to average over local variations among particles.

2.3 Stone characterisation and sols application

The three types of stones investigated were limestones known as Maccabim, Halila and Hebron, respectively. Four
samples for each stone, with average dimensions of ∼ 3.0 × 3.0 × 1.0 cm3 and weight in the range 20–25 g, were
investigated to establish a comparison between the different TiO2 treatments on the stones and the untreated stone
substrates. A first analysis was made on petrographic thin sections by using a Nikon Eclipse E 400 POL polarizing
optical microscope. X-ray diffraction powder patterns were collected at room temperature in 2θ range of 10◦ –70◦ by
Bruker D2 PHASER theta/theta diffractometer, equipped with a one-dimensional “compound silicon strip” detector,
using Cu Kα radiation at 30 kV and 10 mA, at 0.02◦ step size, count time 1 s. Measurements were performed on
samples spread on a zero background sample holder. The reflections of powdered crystalline silicon were used for 2θ
calibration. XRD patterns were compared with standard calcite and dolomite (JCPDS cards #05–0586 and #05–0622,
respectively). The Raman spectra of the stones were acquired using the same Raman apparatus described before for the
TiO2 powder characterisation, in a configuration comprising of a continuous-wave single frequency Nd:YAG blue laser
at 473.1 nm end-pumped by a laser diode. The laser beam was focused on the sample with a spot diameter of nearly 1 μm
(×50, NA = 0.55), using a confocal hole of 150 μm; the spectral resolution was about 2.5 cm−1 . Analyses of the stone
surfaces were performed with a Scanning Electron Microscope (Jeol JSM 6400) equipped with an Oxford Instruments
Link Analytical Si(Li) Energy Dispersive System detector (SEM-EDS). The data analysis was performed by INCA
built-in software, taking calcium as reference. The porosity and the pore size distributions of the untreated stones were
evaluated by Mercury Intrusion Porosimetry analysis using a PoreMaster-33 device from Quantachrome Instruments.
The titania coatings were applied on the stones by a brush. Before treatment, the specimens were washed with
distilled water and dried to constant weight. Aqueous suspension (0.1 M) was brushed on the stones: N-TiF sol was
applied as prepared, while N-TiA was repeatedly washed and re-suspended in distilled water to reach a final pH ≈ 5.
Treatment with TiO2 P25 (Evonik) was used as a comparison, applying a 0.1 M aqueous dispersion (pH ≈ 5). To
guarantee the whole surface covered, three layers of TiO2 nano-sol were applied. Then, the stones were kept drying
in a ventilated oven at 50 ◦ C for 60 min. The amount of coating applied per unit surface was estimated from the net
quantity of TiO2 sol used to coat the stones. Values are averaged on four samples treated with each coating.
The compatibility of the coatings was tested, according to UNI-EN standards, by measurements of the colour
variations and by capillary water absorption tests. Colorimetric coordinates were measured in the CIEL∗ a∗ b∗ system
by a Techkon Spectrodens colorimeter with D65 illuminant (average daylight and correlated colour temperature of
approximately 6500 K) and 10◦ viewing angle geometry, wavelength range between 400 and 700 nm [20]. The colour
changes of the surfaces induced by the TiO2 treatments were evaluated according to the norm UNI EN 15886:2010 [21].
Measurements were collected on twelve round regions of about 3 mm diameter on each sample and averaged on four
samples. Colour differences were measured in the CIEL∗ a∗ b∗ standard colour system, evaluated by

ΔE ∗ = ΔL∗2 + Δa∗2 + Δb∗2 , (1)

where ΔL∗ , Δa∗ and Δb∗ are the colorimetric coordinate differences before and after the coating.
Page 4 of 13 Eur. Phys. J. Plus (2019) 134: 539

The capillary water absorption was performed according to the norm UNI EN 15801:2010 [22]. Two treated samples
for each TiO2 coating type were employed and compared with the corresponding untreated stones. The amount of
absorbed water per unit area Qi (kg/m2 ) at time ti was calculated by Qi = (mi −m A
0)
, m0 being the mass of the dry
specimen, mi the mass at time ti and A the stone surface in contact with the wet paper. The Qi values were reported as
1/2
a function of the square root of time (ti ), according to the rules. The capillary water absorption coefficient (kg/(m2
1/2
s )) was determined by the slope of the curve calculated by linear regression, using at least 5 consecutive points at
low times (t < 60 min) [23].

2.4 Photoinduced properties

Photoinduced hydrophilicity and photocatalytic activity were investigated by exposing the TiO2 coated surfaces to UV-
Vis light. Untreated specimens served as references. The wettability of the stone surfaces was evaluated by static contact
angle (CA) measurements, according to the UNI EN 15802:2010 [24], performed by OCA 20 apparatus (DataPhysics
Instrument GmbH) using the sessile drop method, with 1 μL drops applied on the surface by a needle. Contact angles
were evaluated also on samples irradiated by an Osram Ultra-Vitalux (300 W) UV-Vis lamp placed 20 cm from the
samples, with a UV irradiance power of about 1 mW/cm2 . Contact angles were obtained from the intercept at t = 0 of
a linear fit to the contact angle values measured every 0.1 s for 50 s after the deposition, by recording the drop shape
with a camera. CA measurements were repeated at increasing irradiation times (1 – 3 – 5 – 8 min).
The photocatalytic activity of the synthesised N-doped TiO2 on the stones was tested choosing as pollutant agents
two synthetic organic dyes, Rhodamine B and Methyl orange. N-doped TiO2 coated stone specimens were stained
by brush with three layers of an aqueous solution of the organic dye (1 mM) and then kept to dry in a ventilated
oven at 50 ◦ C for 60 min. The photocatalytic test was carried out under the UV-Vis Osram Ultra Vitalux lamp in the
same irradiation conditions as for the CA measurements. The photo-degradation activity on each dye was evaluated
by colorimetric measurements averaged on twelve circular regions (3 mm diameter) for each stained sample surface
on two samples treated with the same coating, using the Techkon Spectrodens colorimeter. Colour variations were
recorded over time: a normalised chroma change ΔCn at different exposure time t was obtained by the chromatic
coordinates a∗ (t) and b∗ (t): 
(a∗ (t) − a∗ (0))2 + (b∗ (t) − b∗ (0))2
ΔCn = , (2)
(a∗C − a∗ (0))2 + (b∗C − b∗ (0))2
where a∗ (t) and b∗ (t) are the coordinates at irradiation time t, while a∗C and b∗C are measured on the clean stones before
staining with dyes [25,26]. This function should reach 1 after the complete dye degradation to uncoloured species.

3 Results and discussion


3.1 TiO2 powder characterisation

XRD patterns of the new neutral N-TiF and of the acidic N-TiA nanopowders are shown in fig. 1. Both samples
exhibit a nanocrystalline structure: the observed main diffraction peaks can be indexed by crystallographic planes of
anatase (marked with A), while the small diffraction peak observed at 2θ = 30.81◦ (marked with B) in N-TiA can
be ascribed to a brookite phase. The weight fraction of anatase and brookite phases were estimated, as described
in [17,27], for N-TiA as 0.80 ± 0.03 and 0.20 ± 0.04, respectively (see inset in fig. 1), while for N-TiF the brookite
contribution was too low to be determined. XRD measurements on Aeroxide P25 revealed a rutile weight fraction of
0.21 ± 0.02, like commonly reported values [28], and absence of any brookite phase. The weight percentage of the TiO2
phases in each product is summarised in table S1 in the Supplementary Material (SM). By the FWHM of the main
diffraction peak (101) of anatase, the Scherrer formula in the spherical approximation gives the average crystallite size
of the nanoparticles: 4.5 ± 1.0 nm and 3.5 ± 1.0 nm in N-TiF and N-TiA, respectively. For P25 a rough estimate of the
nanocrystals size gives 17 ± 4 nm for anatase and 25 ± 4 nm for rutile [28].
N-TiF and N-TiA samples, diluted 1:100 and 1:200 in ethanol, respectively, were investigated by TEM. High-
Resolution TEM (HRTEM) images (fig. 2) show interplanar spacing of 0.35 nm and 0.24 nm, corresponding to the
(101) and (004) planes of anatase, respectively [29]. N-TiA nanoparticles (fig. 2(a)) show a nearly spherical shape with
sizes in the 3–5 nm range. In the case of N-TiF nanoparticles (fig. 2(b)), the morphology of the aggregates suggests a
predominant rod-shaped structure, with average 4 × 9 nm2 cross sections.
The Raman spectra of N-doped TiO2 nanopowders, N-TiF and N-TiA, and of P25 are reported in fig. 3, compared
to the crystalline anatase spectrum, which shows the Eg , Eg , B1g , A1g (B1g ), and Eg modes at 143, 197, 395, 515,
638 cm−1 , respectively. The structure of the Raman spectrum of P25 is similar to that of crystalline TiO2 (peaks at
145, 199, 398, 520, 640 cm−1 ) but shows in addition weak features due to the rutile phase. For the N-doped powders,
Eur. Phys. J. Plus (2019) 134: 539 Page 5 of 13

Fig. 1. XRD patterns of N-TiA and N-TiF powders: (A) anatase, (B) brookite. In the inset: fit of the main anatase-brookite
peaks.

Fig. 2. HRTEM images of N-TiA (a) and N-TiF (b) nanoparticles: the crystalline planes of anatase may be identified.

the five high-intensity Raman peaks, at 152–154 cm−1 , 204 cm−1 , 398 cm−1 , 512 cm−1 , and 635 cm−1 , correspond to
the characteristic Raman modes of the TiO2 anatase structure. The main peak is therefore shifted from 143 cm−1 to
about 152–154 cm−1 and its width (FWHM) is more than doubled (from 12 cm−1 to ≈ 30 cm−1 ). The peaks at higher
wavenumbers are broader and slightly shifted compared to crystalline anatase [30]. This effect is usually interpreted as
due in large part to a phonon quantum confinement effect which occurs when the particle size is less than 20–30 nm,
even if deviation from stoichiometry or the presence of defects may play an important role [30,31]. The confinement
effect is indeed nearly negligible for P25, with nanocrystalline sizes of about 20 nm. The broadening of the main
anatase peak (FWHM: 12–14 cm−1 in crystalline TiO2 and in P25, 28 cm−1 in N-TiF and 33 cm−1 in N-TiA) and the
blueshift up to 154 cm−1 observed in the synthesised samples, are compatible with anatase nanocrystals of 3–6 nm
size, also taking into account the brookite contribution. The features from the brookite phase in N-TiA are too weak
to be observed: the main brookite peak at 154 cm−1 overlaps with the largest anatase peak.
Page 6 of 13 Eur. Phys. J. Plus (2019) 134: 539

Fig. 3. Raman spectra of TiO2 nanopowders. The wavenumber of the main anatase peaks (A) are shown. R indicates features
due to the rutile phase.

Fig. 4. UV-Vis absorption spectra of N-TiA, N-TiF, the reference crystalline anatase TiO2 and P25. (a) log(1/R) versus λ
(R is the reflectance); (b) Tauc plot of (F (R)hν)1/2 versus photon energy and linear fits of the curves. Calculated band gap
energies (in eV) are indicated.

Figure 4(a) shows the absorbance log(1/R) of the various types of N-TiO2 powders versus the wavelength λ,
as measured by DRS. Based on these spectra, the energy bandgap Eg values of the various types of powders were
estimated using a Kubelka-Munk (KM) function. The absorption coefficient K and the scattering coefficient S are
related to the KM function F (R) = (1 − R)2 /2R = K/S [18]. As S is roughly independent of the wavelength, the
KM function may be taken proportional to the absorption coefficient and Eg may be estimated by the Tauc plot
(hνK)n ∝ (F (R)hν)n ∝ (hν − Eg ), by a fitting of the linear part of the curve, by extrapolating to F (R) = 0.
n = 2 should indicate direct allowed transitions and n = 1/2 indirect allowed transitions: an indirect band gap is
usually reported for anatase. Tauc plots of (F (R)hν)1/2 versus hν for N-TiA and N-TiF, the references P25 and
crystalline anatase and the linear fits are shown in fig. 4(b). Both N-doped TiO2 nanopowders show a red-shift in
their absorption spectra compared to P25 and to crystalline anatase TiO2 (3.07 and 3.17 eV for N-TiA and N-TiF,
respectively). The different behavior of P25 from crystalline anatase is related to the rutile (∼ 20%) content of this
commercial product [18]. The changes in the optical absorption in the N-doped samples are ascribed to the introduction
of localised N2p doping states in the band gap of TiO2 near to the valence band edge, due to substitutional and/or
interstitial nitrogen. The absorption enhancement in the visible-light range corresponds to a very slight yellowish
colour of the powders, noticeable in particular for the N-TiA samples.
XPS measurements performed on both N-TiA and N-TiF powders (fig. 5) provide important information on
the two powders. The atomic ratio between O:Ti:N:F is found to be 3.2(±0.2):1:0.22(±0.1):0 for N-TiA and
4.5(±0.4):1:0.15(±0.07):0.15(±0.02) for N-TiF. It is worth noticing the high O:Ti ratio and the presence of signif-
icant amount of fluorine in N-TiF. The Ti2p doublet (2p3/2 , 2p1/2 ), representing the Ti4+ species in pristine TiO2 is
Eur. Phys. J. Plus (2019) 134: 539 Page 7 of 13

Fig. 5. XPS spectra of N-doped TiO2 . (a) Ti2p peak of N-TiA; (b) N1s peak of N-TiA; (c) Ti2p peak of N-TiF; (d) N1s peak
of N-TiF.

known to have electron binding energies of 458.5 ± 0.4 eV and 464.5 ± 0.4 eV, respectively [32]. These binding ener-
gies are shifted to two different directions in the nitrogen-doped samples. For N-TiA, the 2p3/2 has another peak at
456.7 eV, i.e. at binding energy lower by approximately 1.5 eV than that of Ti4+ species in pristine TiO2 . As can be
expected, the Ti2p1/2 also reveals a second contribution at lower binding energy (fig. 5(a)). In contrast, the Ti2p3/2
in the N-TiF powder (fig. 5(c)) is shifted to higher binding energies, showing another peak at 460.6 eV, together with
very wide envelope peaked at 463.8 eV and at 467.6 eV that reveals a long tail. The contradictory effects of the two
N-doping preparation techniques (N-TiA and N-TiF) on the Ti XPS signal suggest that the doping procedure may
have a significant effect on the local environment of the titanium atoms, which eventually may affect the photocatalytic
activity. It should be noted that the Ti peak in TiN was reported to be located around 455 eV [33], hence, it is unlikely
that the shift towards higher binding energies in N-TiF was due to the presence of nearby nitrogen atoms. Such a
shift is reasonable if the titanium atoms are bonded to highly electronegative fluorine atoms, as found in significant
concentration in N-TiF. Indeed, the Ti2p1/2 peak in TiF4 was reported to be located at 466.9 eV [34]. The nitrogen
N1s signal in N-TiA is located at 400.2 eV (fig. 5(b)), whereas for N-TiF (fig. 5(d)) there is one peak at 401.6 eV and
a second one at 404.9 eV, i.e. at binding energy higher by 4 eV. Previous work asserted that the 1 s electron binding
energy for interstitial nitrogen in titania should be around 400–401 eV and that, for interstitial and substitutional
nitrogen, the two peaks should be well resolved [35]. To some extent, this is in contrast with published data on TiN,
where the N1s peak was at 397 eV [33] and is shifted to lower energies by 0.5–1 eV in the presence of oxygen, i.e. in
TiOx Ny [36]. The location of the nitrogen peak cannot be taken, therefore, as indicative for substitutional nitrogen
in both N-TiA and N-TiF types of samples. In this context it should be noted that the exact origin of the interstitial
peak of nitrogen in titanium oxide is not clear and was hypothesised as N2 , NO3 − or N-H [37], which seems to be in
line with the finding of a 408 eV peak in some of the N-TiA samples. An interesting XPS result is the considerably
higher than 2 ratio between the atomic concentration of oxygen to that of titanium, both in N-TiA (3:1) and in N-TiF
(4:1). This may be originated from the presence of adsorbed oxygen-containing molecules at the surface. However, in
that case, one could expect the oxygen/titanium ratio in the surface mode of XPS measurements to be higher than in
the bulk mode, in contrast to our observations that showed almost no difference between the two modes of operation.
On the other hand, if (part) of the nitrogen appears as nitrate, one may expect the atomic percentage of oxygen to
be higher than that of stoichiometric TiO2 .
Page 8 of 13 Eur. Phys. J. Plus (2019) 134: 539

Fig. 6. Stone samples (in the insets) and optical microscope images of thin sections: (a) Maccabim (b) Halila and (c) Hebron.

Fig. 7. XRD of Maccabim, Halila and Hebron stones with dolomite and calcite peaks indexing.

3.2 Stone characterisation and sols application

Maccabim stone shows a heterogeneous pink colour with white-yellow creamish sections, while Halila and Hebron
stones have a more uniform colour, yellowish and white, respectively. Some red stylolites are characteristic of Hebron
and, to a lesser extent, of Maccabim stones. From the optical microscope images, the stones can be distinguished by
the micritic structure of Maccabim (fig. 6(a)), while in Halila and Hebron stones calcite clasts can be observed among
the cementing material (figs. 6(b)–(c)). Colorimetric measurements of the investigated stones are reported in table S2
(SM). XRD patterns obtained on powdered samples of the stones (fig. 7) attest the prevailing presence of dolomite
in Maccabim stones, with a residual content of calcite. On the contrary, Halila stone consists mainly of calcite with a
very low percentage of dolomite. Hebron stone matches calcite diffraction pattern. Raman spectroscopy and SEM-EDS
confirmed the dolomitic and calcitic nature of the stones and evidenced goethite as a minority phase, as reported in
figs. S1–S2–S3 (SM). Porosimetric analysis evidences the highest porosity in Maccabim stones (9.1 ± 0.1)%, while total
porosity in Halila and Hebron are quantified at and 5.4% and 4.9%. Median pore size is evaluated 0.36 μm, 0.25 μm
and 0.23 μm (±0.01 μm) for Maccabim, Halila and Hebron stones, respectively. The pore size distribution is reported
in fig. S4 (SM).
The amount of TiO2 nano-sols applied to the stone surfaces was measured and kept constant, as accurately as
possible, for all coatings (detailed values are summarised in table 1). The distribution of the TiO2 coatings was assessed
by SEM-EDS measurements: examples of surface and depth distributions of TiO2 are reported in figs. S5–S6–S7 (SM).
The differences in the quantity of applied product can be ascribed to the different homogeneity of the coatings.
The synthesised N-TiO2 are aqueous suspensions while the P25 based coating is an aqueous dispersion, making its
application difficult and inhomogeneous due to the agglomeration of TiO2 nanoparticles.
Eur. Phys. J. Plus (2019) 134: 539 Page 9 of 13

Table 1. Amount of TiO2 applied product on the stone samples (SD are reported in brackets).

Amount of applied product [g/m2 ]


N-TiA N-TiF P25
Maccabim 1.7(0.4) 1.4(0.2) 1.1(0.1)
Halila 1.5(0.1) 1.2(0.1) 1.0(0.1)
Hebron 1.2(0.1) 0.9(0.2) 0.8(0.2)

Fig. 8. Capillarity water absorption tests carried on (a) Maccabim, (b) Halila and (c) Hebron stones samples. N-TiA, N-TiF
and P25 coated stones are compared with the untreated ones.

Table 2. Capillarity water absorption coefficient (AC) on Maccabim, Halila and Hebron stones. N-TiA, N-TiF and P25 treated
stones are compared to the untreated ones.

AC (kg m−2 s−1/2 ) × 10−3


Untreated N-TiA N-TiF P25
Maccabim 9.7 8.4 9.2 8.6
Halila 3.2 3.3 3.7 3.2
Hebron 2.4 1.9 3.1 1.5

The amount of water absorbed by the specimens per unit area Qi (kg/m2 ) at time ti is shown in fig. 8 as a function
1/2
of the square root of time (ti ). Results show that there are few significant differences in the trend before and after the
application of the coating. The Qi values change among the different types of stones mainly because of the difference
in the substrate texture rather than due to the TiO2 treatments. Maccabim stone (fig. 8(a)), due to its higher porosity,
shows higher Qi values than Halila (fig. 8(b)) and Hebron (fig. 8(c)) stones, more compact. After all treatments, on
Maccabim stone there is a slight increase (5–7%) in the water absorption, while N-TiA and P25 treatments on Hebron
stone induce a slight reduction. Capillarity absorption coefficients (AC), shown in table 2, were obtained by linearly
fitting the Qi data points taken during the first hour of measurements vs. t1/2 . The trend of the Qi curves during the
whole duration of the capillarity tests is confirmed by the AC values that are almost the same for the treated and the
untreated stones.
Page 10 of 13 Eur. Phys. J. Plus (2019) 134: 539

Table 3. Colorimetric variations induced on the stones by TiO2 coatings, expressed by ΔE ∗ in the CIEL∗ a∗ b∗ system.

N-TiA N-TiF P25


Maccabim 2.4 ± 0.3 1.2 ± 0.1 1.4 ± 0.1
Halila 3.3 ± 0.2 1.8 ± 0.2 1.7 ± 0.1
Hebron 2.6 ± 0.1 1.0 ± 0.1 1.1 ± 0.1

Fig. 9. Contact angles as a function of the UV-Vis irradiation time. An average standard error of about 3◦ may be assumed.

In table 3 the colour variations, due to the TiO2 coatings, are expressed through the total colour difference ΔE ∗ ,
according to the UNI EN 15886:2010 [21]. The N-TiA coating gives the highest colour variations, but close to the
limit of perceptibility, especially on Halila stone. The colour difference induced by the N-TiF is lower, comparable
with that found for P25. Various acceptability criteria suggest ΔE ∗ slightly greater than the limit of perceptibility by
human eye which is considered in the range 2–3 [38]. Often, for Cultural Heritage applications, a less restrictive limit
ΔE ∗ < 5 is accepted [39].

3.3 Photoinduced properties

The photoinduced hydrophilicity of N-doped TiO2 coatings on the three different stones was investigated by surface
wettability effects, measuring the static contact angle (CA) on treated and untreated stones, before and after irradiation
with UV-Vis light. It should be stressed that the determination of the contact angle is quite difficult: the results are
strongly influenced by the different morphology and physical characteristics (porosity, roughness) of the stone samples.
Before the irradiation, significant differences between untreated and treated stones are clearly visible (fig. 9). On all
the different types of stones, only the N-TiA treatment produces an increase in the contact angle compared to the
untreated surfaces, with a ∼ 40%, ∼ 55% and ∼ 90% increase on Maccabim, Halila and Hebron stones, respectively.
This effect may be attributed to the higher homogeneity of the N-TiA coating [17]. Following the irradiation with
UV-Vis light, both N-TiO2 coatings show a strong decrease in the contact angle. N-TiA coating induces the most
evident hydrophilicity, reaching the lowest CA values in the shortest irradiation times: for times longer than 8 minutes
the contact angle for N-TiA is far less 10◦ . However, the photoinduced hydrophilicity in N-TiF coated samples is
higher than in P25 coated ones. An example of the change in the shape of the water droplets, on samples coated with
N-TiF, in comparison with the untreated stone, is shown in fig. 10.
Eur. Phys. J. Plus (2019) 134: 539 Page 11 of 13

Fig. 10. Water droplets over (a) untreated Maccabim stone sample, (b) N-TiF coated stone before irradiation, (c) N-TiF coated
stone after 5 minutes of UV-Vis irradiation.

Fig. 11. Chroma variations ΔCn as a function of irradiation time for RhB-stained samples: (a) Maccabim, (b) Halila and (c)
Hebron stones, untreated and treated with N-TiA, N-TiF and P25 as a reference.

The photocatalytic activity of the synthesised N-TiO2 nano-sols was then tested and compared on the three types
of stones. The stones were treated with N-TiF and N-TiA coatings are then stained with two different dyes, Rhodamine
B (RhB) and Methyl orange (MeO). The photodegradation of the dyes under UV-Visible illumination was investigated
through colorimetric measurements at given time intervals. Change in the chroma ΔCn (eq. (2)) was chosen to monitor
the colour variation as a function of the irradiation time. The discolouration of RhB on the three types of stones is
shown in fig. 11, where the normalised ΔCn is reported as a function of the irradiation time. The photocatalytic
efficiency of N-TiF coating is compared with those of N-TiA and P25. Both N-doped TiO2 coatings exhibit good
discolouring performance: in the first 20 minutes the degradation of the RhB dye reaches 20–30%, up to more than
60–70% in about 200 min.
Due to the self-fading of the dye, partial discolouration of untreated stones is observed, amounting to 30–40%. In
general, the visible-light photocatalytic activity of doped TiO2 should be tested with molecules that do not absorb
light at the irradiation wavelength, to prevent both photolysis and dye photo-sensitization effects [40–42]. On the
other hand, for practical assessment it is less important if the colour change involves not only photocatalysis but also
some dye photosensitization. In any case, the standard UNI-EN rules for the photocatalytic activity determination in
building materials still require a RhB based test [43]. The discolouration of MeO on the three limestones is shown in
fig. 12. The highest photodegradation is achieved in all the three stones by N-TiA coating: 90% of dye discolouration
is reached in 210 minutes of irradiation. The performance of N-TiF coating is in any case better than the P25. The
faster discolouration of N-TiA coating of MeO may be explained by the fact that anionic azo-dyes are better adsorbed
when the surface of the TiO2 nanoparticles is positively charged, as in the case of N-TiA [16,17]. As for RhB, a partial
discolouration of untreated stones is observed. Here, the photocatalytic activity of N-TiF coating is even comparable
to N-TiA efficiency and in any case better than the P25.
Page 12 of 13 Eur. Phys. J. Plus (2019) 134: 539

Fig. 12. Chroma variations ΔCn as a function of irradiation time for MeO-stained samples: (a) Maccabim, (b) Halila and (c)
Hebron stones, untreated and treated with N-TiA, N-TiF and P25.

The discolouration kinetics of the three types of stones does not reveal much differences. While this is expected
for smooth surfaces over-coated with optically thick photocatalytic layer, the picture can be more complex when
the surface has micropores, or when the photocatalyst is introduced into a cementitious bulk [44]. In our case, the
difference between the photodegradation kinetics in coated and uncoated Maccabim is higher than in the other stones.
This effect is more noticeable in the degradation of MeO than for RhB. One possible explanation is its higher porosity,
making this stone more penetrable to water. In this case, the dye penetrates deeper into the Maccabim stone, thus
reducing the rate of discolouring by direct photolysis, but not that of photocatalysis, which may operate also in “dark”
microdomains, either by a mechanism involving the movement of adsorbed molecules [45], or by a mechanism involving
the transport of active species [46,47].

4 Conclusions

The acidic and neutral aqueous media allowed the synthesis of nanostructured N-doped TiO2 (N-TiF and N-TiA,
respectively) with crystalline different shape and size (4–10 nm range), as shown by TEM, XRD and Raman analyses.
N-TiA prefers a spherical shape whereas N-TiF nanoparticles are more elongated. The main detected phase is anatase,
with a small brookite content in N-TiA. Band gap evaluation by Tauc plot in Kubelka-Munk approximation made by
DRS measurements suggests a light shift toward lower energies of the N-TiO2 samples rather than the commercial
product P25 by Evonik (nanocrystalline anatase/rutile at 80%/20%) and crystalline TiO2 . XPS measurements of the
powders showed O:Ti atomic ratio considerably larger than stoichiometric (3 for N-TiA and 4 for N-TiF). Analysis
of the shifts of the Ti2p and N1s peaks, even if not conclusive, suggests that nitrogen atoms are not located in
substitutional positions.
Both N-doped TiO2 coatings, applied on dolomitic and calcitic limestones from Israel, exhibit effective photo-
catalytic activity under UV-Visible irradiation, without altering the appearance and the capillary water absorption
properties of the stones. A significant increase in the surface wettability was observed following exposure to UV-Visible
light, for both N-doped coatings. For short times of irradiation (8 mins) the contact angle values approach the super-
hydrophilicity limit for N-TiA. N-TiF and N-TiA show good performances in photocatalytic activity for the Methyl
orange degradation in stained limestone samples, higher or comparable to that of the commercial product P25. On
the other hand, the photocatalytic activity of titania for Rhodamine B is still matter of discussion due to the dye
photosensitization and needs further investigation, especially when tested on stones, a topic where literature is scarce.
The N-TiF coating, even if it shows self-cleaning and photocatalytic performances slightly less satisfactory than
those of N-TiA, thanks to the intrinsic neutral character may be a good choice for limestones used in buildings and
cultural heritage artefacts.
Eur. Phys. J. Plus (2019) 134: 539 Page 13 of 13

Support from MAECI (project “NANO4HER, Nanotechnology at the service of cultural heritage preservation”, Italy-Israel
Scientific and Technological Cooperation, Ministry of Science, CUP number: D82I14000470001) is gratefully acknowledged.

Publisher’s Note The EPJ Publishers remain neutral with regard to jurisdictional claims in published maps and institutional
affiliations.

References
1. P. Munafò, G.B. Goffredo, E. Quagliarini, Constr. Build. Mater. 84, 201 (2015).
2. E. Quagliarini, F. Bondioli, G.B. Goffredo, A. Licciulli, P. Munafò, J. Cult. Herit. 14, 1 (2013).
3. M.J. Mosquera, D.M de los Santos, T. Rivas, P. Sanmartı́n, B. Silva, J. Nano Res. 8, 1 (2009).
4. M. Pelaez, N.T. Nolan, S.C. Pillai, M.K. Seery, P. Falaras, A.G. Kontos, P.S.M. Dunlop, J.W.J. Hamilton, J.A. Byrne, K.
O’Shea, M.H. Entezari, D.D. Dionysiou, Appl. Catal. B - Environ. 125, 331 (2012).
5. A. Sierra-Fernandez, L.S. Gomez-Villalba, M.E. Rabanal, R. Fort, Mater. Construcc. 67, 325 (2017).
6. S. Banerjee, D.D. Dionysiou, S.C. Pillai, Appl. Catal. B - Environ. 176-177, 396 (2015).
7. O. Sacco, V. Vaiano, L. Rizzo, D. Sannino, J. Clean. Prod. 175, 38 (2018).
8. A. Calia, M. Lettieri, M. Masieri, S. Pal, A. Licciulli, V. Arima, J. Clean. Prod. 165, 1036 (2017).
9. M.Z. Guo, A. Maury Ramirez, C.S. Poon, J. Clean. Prod. 112, 3583 (2016).
10. L. Bergamonti, F. Bondioli, I. Alfieri, A. Lorenzi, M. Mattarozzi, G. Predieri, P.P. Lottici, Appl. Phys. A 122:124, 1 (2016).
11. W. Shen, C. Zhang, Q. Li, W. Zhang, L. Cao, J. Ye, J. Clean. Prod. 87, 762 (2015).
12. E. Quagliarini, F. Bondioli, G.B. Goffredo, C. Cordoni, P. Munafò, Constr. Build. Mater. 37, 51 (2012).
13. M.F. La Russa, S.A. Ruffolo, N. Rovella, C.M. Belfiore, A.M. Palermo, M.T. Guzzi, G.M. Crisci, Prog. Org. Coat. 74, 186
(2012).
14. R. Rahal, A. Wankhade, D. Cha, A. Fihri, S. Ould-Chikh, U. Patil, V. Polshettiwar, RSC Adv. 2, 7048 (2012).
15. S.G. Kumar, K.S.R. Koteswara Rao, Appl. Surf. Sci. 391, 124 (2017).
16. L. Bergamonti, I. Alfieri, A. Lorenzi, A. Montenero, G. Predieri, R. Di Maggio, F. Girardi, L. Lazzarini, P.P. Lottici, J.
Sol-Gel Sci. Technol. 73, 91 (2015).
17. L. Bergamonti, G. Predieri, Y. Paz, L. Fornasini, P.P. Lottici, F. Bondioli, Microchem. J. 133, 1 (2017).
18. R. López, R. Gómez, J. Sol-Gel Sci. Technol. 61, 1 (2012).
19. K.M. Reddy, S.V. Manorama, A.R. Reddy, Mater. Chem. Phys. 78, 239 (2003).
20. C. Oleari, Standard Colorimetry: Definitions, Algorithms and Software, 1st ed. (John Wiley & Sons, Chicester, UK, 2016).
21. UNI EN 15866:2010, Conservation of cultural property – test methods – colour measurement of surfaces (UNI Ente Nazionale
Italiano di Unificazione, Milano, 2010).
22. UNI EN 15801:2010, Conservation of cultural property – test methods – Determination of water absorption by capillarity
(UNI Ente Nazionale Italiano di Unificazione, Milano, 2010).
23. E.W. Washburn, Phys. Rev. 17, 273 (1921).
24. UNI EN 15802:2010, Conservation of cultural property – test methods – Determination of static contact angle (UNI Ente
Nazionale Italiano di Unificazione, Milano, 2010).
25. L. Bergamonti, I. Alfieri, A. Lorenzi, A. Montenero, G. Predieri, G. Barone, P. Mazzoleni, S. Pasquale, P.P. Lottici, Appl.
Surf. Sci. 282, 165 (2013).
26. L. Bergamonti, I. Alfieri, M. Franzò, A. Lorenzi, A. Montenero, G. Predieri, M. Raganato, A. Calia, L. Lazzarini, D. Bersani,
P.P. Lottici, Environ. Sci. Pollut. R. 21, 13264 (2014).
27. H. Zhang, J.F. Banfield, J. Phys. Chem. B 104, 3481 (2000).
28. B. Ohtani, O.O. Prieto-Mahaney, D. Li, R. Abe, J. Photochem. Photobiol. A 216, 179 (2010).
29. H.E. Swanson, H.F. McMurdie, M.C. Morris, E.H. Evans, Nat. Bur. Stand. (U.S.) Monogr. 25, 7 (1969).
30. D. Bersani, P.P Lottici, X.Z. Ding, Appl. Phys. Lett. 72, 73 (1998).
31. A. Golubović, M. Šćepanović, A. Kremenović, S. Aškrabić, V. Berec, Z. Dohčević-Mitrović, Z.V. Popović, J. Sol-Gel Sci.
Tecnol. 49, 311 (2009).
32. W.E. Slink, P.B. DeGroot, J. Catal. 68, 423 (1981).
33. P. Prieto, R.E. Irby, J. Vac. Sci. Technol. A 13, 2819 (1995).
34. V.D. Klimov, A.A. Vashman, I.S. Pronin, Zh. Obshch. Khim. 61, 2166 (1991).
35. J. Lynch, C. Giannini, J.K. Cooper, A. Loiudice, I.D. Sharp, R. Buonsanti, J. Phys. Chem. C 119, 7443 (2015).
36. A. Glaser, S. Surnev, F.P. Netzer, N. Fateh, G.A. Fontalvo, C. Mitterer, Surf. Sci. 601, 1153 (2007).
37. O. Diwald, T.L. Thompson, T. Zubkov, E.G. Goralski, S.D. Walck, J.T. Yates, J. Phys. Chem. B 108, 6004 (2004).
38. M. Mahy, L. Eycken, A. Oosterlinck, Color Res. Appl. 19, 105 (1994).
39. H.R. Sasse, R. Snethlage, Sci. Technol. Cult. Herit. 5, 85 (1996).
40. M. Rochkind, S. Pasternak, Y. Paz, Molecules 20, 88 (2015).
41. N. Barbero, D. Vione, Environ. Sci. Technol. 50, 2130 (2016).
42. N. Shaham-Waldmann, Y. Paz, Mat. Sci. Semicon. Proc. 42, 72 (2016).
43. UNI EN 11259:2016, Determination of the photocatalytic activity of hydraulic binders – Rodammina Test Method (UNI Ente
Nazionale Italiano di Unificazione, Milano, 2016).
44. N. Abbas, Y. Paz, J. Adv. Oxid. Technol. 19, 218 (2016).
45. A. Avraham-Shinman, Y. Paz, Israel J. Chem. 46, 33 (2006).
46. H. Haick, Y. Paz, J. Phys. Chem. B 105, 3045 (2001).
47. H. Haick, Y. Paz, ChemPhysChem 4, 617 (2003).

You might also like